26.11.2014 Views

(PHH98 2d-Slide#15 \(8 of 8\)) - Atmospheric Dynamics Group

(PHH98 2d-Slide#15 \(8 of 8\)) - Atmospheric Dynamics Group

(PHH98 2d-Slide#15 \(8 of 8\)) - Atmospheric Dynamics Group

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

BALANCED FLOW:<br />

A PROGNOSTIC EQUATION<br />

(PHH lecture 2)<br />

3-D rotating stratified flow<br />

Consider 3-D fluid with density stratification.<br />

Make hydrostatic approximation:<br />

DOMINANT BALANCE IN VERTICAL MOMENTUM<br />

EQUATION IS BETWEEN<br />

VERTICAL PRESSURE GRADIENT<br />

GRAVITATIONAL FORCE<br />

(Can be argued from<br />

vertical length scales


Thermal wind relation<br />

Geostrophic balance: f × u = − 1 ρ0<br />

∇p<br />

Hydrostatic balance: − ∂p<br />

∂z − ρg = 0<br />

[Note: Boussinesq system with geometric height used<br />

as a vertical co-ordinate unless otherwise stated]<br />

Thermal wind balance:<br />

f × ∂u<br />

∂z = g<br />

ρ0<br />

∇ρ<br />

f<br />

u<br />

heavy (cold)<br />

light (warm)<br />

N.B. Meteorology: maps <strong>of</strong> ‘thickness’ may be used to<br />

infer the change in wind with height across a finite<br />

layer. Satellite observations <strong>of</strong> temperature are<br />

used to infer stratospheric winds.<br />

Oceanography: observations <strong>of</strong> ρ are sometimes<br />

used to infer velocities – requires ‘level <strong>of</strong> no motion’<br />

Quasi-geostrophic flow<br />

PHH 2 / 3<br />

We seek a predictive equation for the time evolution <strong>of</strong> a<br />

flow that is always close to geostrophic balance<br />

Scale analysis <strong>of</strong> (hydrostatic) equations<br />

∂u<br />

∂t<br />

+ u.∇u + f × u = − ∇p<br />

ρ0<br />

U<br />

T<br />

1<br />

f T<br />

U 2<br />

L<br />

U<br />

f L<br />

fU<br />

1 relative sizes<br />

‘ Rossby number’ Ro<br />

1<br />

f T or U f L<br />

Small Rossby number requires<br />

time scale > ≈ few days (at midlatitudes) and<br />

U


• Geostrophic balance gives us no predictive equation<br />

for the motion (at least if f is constant) because mass<br />

continuity follows identically from the geostrophic<br />

flow<br />

• Need to consider the small departures from<br />

geostrophic balance<br />

Effects <strong>of</strong> spherical geometry:<br />

• For large scale motion in the atmosphere and the<br />

ocean we need to take account <strong>of</strong> some <strong>of</strong> the effects <strong>of</strong><br />

sphericity<br />

• Rossby’s great insight was that the most important<br />

effect is the latitudinal variation <strong>of</strong> the Coriolis<br />

parameter:<br />

y latitude co-ordinate<br />

f = f0 + βy f0 , β constants PHH 2 / 5<br />

[β-plane approximation]<br />

f0 = 2Ωsin φ 0<br />

β = 2Ω a cos φ 0<br />

⎫<br />

⎪<br />

⎬<br />

⎪<br />

⎭<br />

where φ0<br />

is latitude<br />

Derivation <strong>of</strong> quasi-geostrophic equations<br />

[For careful mathematical derivation see Pedlosky<br />

(Chapter 6)]<br />

Consider primitive equations on a β-plane, and write<br />

ρ = ρ ′ + ρb( z), p = p ′ + pb(z), where ρb is a ‘ basic state’<br />

stratification and pb is the associated pressure field in<br />

hydrostatic balance<br />

∂<br />

∂t u h + u.∇u h + (f 0 + βy)k × u h = − 1 ρ0<br />

∇h p ′<br />

∇.u = 0<br />

⎛<br />

⎜<br />

⎝<br />

u h = (u,v,0)<br />

u = (u,v, w)<br />

⎞<br />

⎟<br />

⎠<br />

− ∂ p ′<br />

∂z − ρ ′ g = 0<br />

∂ ρ ′<br />

∂t + (u.∇) ρ ′ + w ∂ρ b<br />

∂z = 0<br />

Divide velocity into geostrophic and ageostrophic parts<br />

u = u g + u a<br />

f0 k × u g = − 1<br />

ρ0<br />

∇ p ′<br />

(defines geostrophic velocity)<br />

w = wg + w a , but ∇.u = 0 ⇒ ∂w g<br />

∂z = 0 ⇒ w g = 0<br />

PHH 2 / 6


Ageostrophic part <strong>of</strong> equations<br />

∂<br />

∂t (u g + u a ) h + (u g + u a ).∇(u g + u a ) h<br />

+ f0k × ua + βyk × (u g + u a) = 0<br />

∇.u a = 0<br />

∂ ρ ′<br />

∂t + (u g + u a ).∇ ρ ′ + w a<br />

dρb<br />

dz = 0<br />

− ∂ p ′<br />

∂z − ′<br />

This term retained – implicit<br />

assumption that basic state vertical<br />

density gradient is strong<br />

From vorticity equation<br />

ρ g = 0<br />

using continuity equation for u a,<br />

∂<br />

∂t ζ g + u g .∇ζ g + βv g = − f 0<br />

⎧<br />

⎨<br />

⎩<br />

∂ua<br />

∂x + ∂v a<br />

∂y<br />

⎫<br />

⎬<br />

⎭<br />

= − f0<br />

∂<br />

∂z<br />

⎧<br />

⎪<br />

⎨<br />

⎩ ⎪<br />

1<br />

g ∂ρ b<br />

∂z<br />

⎡<br />

⎢<br />

⎣<br />

∂<br />

∂t<br />

⎛<br />

⎜<br />

⎝<br />

∂ p ′<br />

∂z<br />

⎞<br />

⎟ + u g<br />

⎠<br />

.∇ ⎛<br />

⎜<br />

∂ p ′<br />

⎝ ∂z<br />

⎞<br />

⎟<br />

⎠<br />

⎤<br />

⎥<br />

⎦<br />

⎫<br />

⎪<br />

⎬<br />

⎭ ⎪<br />

then density equation<br />

PHH 2 / 7<br />

Define ψ = p ′/ f0 ρ0<br />

(streamfunction for<br />

quasi-geostrophic motion)<br />

so ug = − ∂ψ<br />

∂y , v g = ∂ψ<br />

∂x , ρ ′ = − f 0ρ0<br />

g<br />

∂ψ<br />

∂z<br />

It follows that<br />

⎧<br />

⎨<br />

⎩<br />

∂<br />

∂t + u g.∇<br />

⎫<br />

⎬<br />

⎭ q g = 0<br />

where<br />

N 2 ( z) = − g<br />

ρ0<br />

dρb<br />

dz<br />

qg = ∂2 ψ<br />

∂x 2 + ∂2 ψ<br />

∂y 2 + ∂ ∂z<br />

⎛<br />

⎜<br />

⎝<br />

2<br />

f0<br />

N 2 ( z)<br />

∂ψ<br />

∂z<br />

⎞<br />

⎟<br />

⎠<br />

+ βy<br />

quasi-geostrophic potential vorticity<br />

Quasi-geostrophic p.v. is conserved following the<br />

geostrophic motion (which is horizontal)<br />

[Rossby–Ertel PV is conserved following the exact motion<br />

– can be shown quite independently <strong>of</strong> any small Ro<br />

assumption (see later).]<br />

PHH 2 / 8


Boundary conditions for<br />

quasi-geostrophic flow<br />

Side boundaries: no normal flow<br />

Top and bottom boundaries: vertical velocity zero (or set<br />

by topography or ‘Ekman pumping’) – deduce from<br />

density equation<br />

⎧<br />

⎨<br />

⎩<br />

∂<br />

∂t + u g . ∇<br />

⎫<br />

⎬<br />

⎭ ρ ′ = −w dρ b<br />

dz<br />

hence<br />

known<br />

⎧<br />

⎨<br />

⎩<br />

∂<br />

∂t + u g . ∇<br />

⎫<br />

⎬<br />

⎭<br />

∂ψ<br />

∂z = − N 2 w<br />

f0<br />

on top or<br />

bottom<br />

boundaries<br />

Top and bottom boundary conditions take the form <strong>of</strong><br />

prognostic equations for surface density or temperature<br />

Similarity to two-dimensional flow:<br />

In 2-D incompressible flow the component <strong>of</strong> vorticity<br />

perpendicular to the plane <strong>of</strong> motion, q, is conserved<br />

and, in terms <strong>of</strong> a streamfunction ψ, the vorticity<br />

equation takes the form<br />

PHH 2 / 9<br />

⎧<br />

⎨<br />

⎩<br />

∂<br />

∂t + u.∇<br />

⎫<br />

⎬<br />

⎭ q = 0<br />

where q = ∂2 ψ<br />

∂x 2 + ∂2 ψ<br />

∂y 2 + βy<br />

⎛<br />

and u = (u, v) = ⎜ ⎝ − ∂ψ<br />

∂y , ∂ψ<br />

∂x<br />

⎞<br />

⎟ .<br />

⎠<br />

[Note that in non-GEFD applications, e.g. aerodynamics,<br />

β would be taken to be zero]<br />

A good deal <strong>of</strong> insight into the behaviour <strong>of</strong> solutions <strong>of</strong><br />

the q.-g. p.v. equation may therefore be gained from the<br />

theory <strong>of</strong> 2-D vortex dynamics (MEM lectures and<br />

various computer demonstrations).<br />

PHH 2 / 10


Different contributions to q g<br />

(and its horizontal gradients)<br />

3-D flow:<br />

qg = ∂2 ψ<br />

∂x 2 + ∂2 ψ<br />

∂y 2 + ∂ ∂z<br />

⎛<br />

⎜<br />

⎝<br />

f0 2<br />

N 2 ( z)<br />

∂ψ<br />

∂z<br />

⎞<br />

⎟<br />

⎠<br />

+ βy<br />

1 2 3<br />

1 Relative vorticity<br />

2 Vortex stretching term (associated with<br />

vertical density gradient)<br />

3 Planetary vorticity<br />

1 and 2 are comparable if f0 L ≈ ND<br />

(Prandtl’s ratio <strong>of</strong> scales)<br />

D<br />

L<br />

( f0 / N < ≈<br />

10−2 for both<br />

atmosphere and ocean,<br />

i.e. thermocline)<br />

1 and 3 are comparable if U ≈ βL2<br />

βL 2 ≈ 10 −1 ms −1 if L ≈ 100 km<br />

10 ms −1 if L ≈ 1000 km<br />

PHH 2 / 11<br />

Numerical simulations <strong>of</strong> geostrophic turbulence<br />

(McWilliams, JFM 1989)<br />

t = 10<br />

t = 20<br />

PHH 2 / 12


Numerical simulations <strong>of</strong> geostrophic turbulence<br />

(McWilliams, JFM 1989)<br />

Isovorticity surfaces cyclones light<br />

anticyclones dark<br />

f/N isotropy with small deviations (e.g. due to fact that<br />

small vertical scales have little relative vorticity)<br />

PHH 2 / 13<br />

Understanding evolution <strong>of</strong> QG turbulence is an ongoing<br />

research problem<br />

(McWilliams et al 1994) find 'columns'as long-time state<br />

A is initial condition, time increases through B,C,D<br />

but see Dritchel & Ambaum (1997) PHH 2 / 14


Summary<br />

The quasi-geostrophic equations are a convenient set<br />

to describe motion at small Rossby number which<br />

remains close to geostrophic balance. [But they are<br />

not quantitatively accurate for modelling realistic<br />

atmosphere or oceanic flow.]<br />

Evolution equation Invertibility principle<br />

qg (quasi - geostrophic<br />

p.v.)<br />

qg and ∂ψ<br />

∂z<br />

(boundaries)<br />

(non-local<br />

operator)<br />

∂ψ<br />

∂z<br />

(temperature at<br />

boundaries)<br />

flow, temperature etc.<br />

in geostrophic balance<br />

GENERALIZATION (see Hoskins et al. 1985)<br />

P (Rossby–Ertel p.v.)<br />

P and θ (boundary)<br />

or σθ (boundary)<br />

θ or σθ (potential temperature<br />

or potential density<br />

at boundaries)<br />

flow, temperature etc.<br />

in balance<br />

(non-local<br />

operator)<br />

PHH 2 / 15

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!